<<

Organozinc compound

Organozinc compounds in organic contain to chemical bonds. Organozinc chemistry is the science of organozinc compounds describing their physical properties, synthesis and reactions.[1][2][3][4]

Organozinc compounds were among the first organometallic compounds made. They are less reactive than many other analogous organometallic reagents, such as Grignard and organolithium reagents. In 1848 prepared the first organozinc Organozinc chemistry compound, , by heating ethyl iodide in the presence of zinc metal.[5] This reaction produced a volatile colorless liquid that spontaneous combusted upon contact with air. Due to their pyrophoric nature, organozinc compounds are generally prepared using air- free techniques. They are unstable toward protic solvents. For many purposes they are prepared in situ, not isolated, but many have been isolated as pure substances and thoroughly characterized.[6]

Organozincs can be categorized according to the number of carbon that are bound to the metal.[2][3]

1. Diorganozinc (R2Zn): A class of organozinc compounds in which two alkyl . These may be further divided into subclasses depending on the other ligands attached 2. Heteroleptic (RZnX): Compounds which an electronegative or monoanionic (X), such as a , is attached to the zinc center with another alkyl or aryl (R). − + 3. Ionic organozinc compounds: This class is divided into organozincates (RnZn ) and organozinc cations (RZnLn ).

Contents Bonding Synthesis From zinc metal Functional group exchange β-Silyl diorganozinc compounds Transmetallation Reactions Reformatsky reaction Simmons–Smith reaction Titanium–zinc methylenation Fukuyama coupling Zinc Organozincates Synthesis Reactions Organozinc(I) compounds See also References External links

Bonding

In its coordination complexes zinc(II) adopts several coordination geometries, commonly octahedral, tetrahedral, and various pentacoordinate geometries. These structural flexibility can be attributed to zinc's electronic configuration [Ar]3d104s2. The 3d orbital is filled, and therefore, ligand field effects are nonexistent. Coordination geometry is thus determined largely by electrostatic and steric interactions.[2] Organozinc compounds usually are two- or three-coordinate, reflecting the strongly donating property of the carbanionic ligands.

[7] Typical diorganozinc complexes have the formula R2Zn. Dialkylzinc compounds are monomeric with a linear coordination at the zinc atom. A polar covalent bond exists between carbon and zinc, being polarized toward carbon due to the differences in values (carbon: 2.5 & zinc: 1.65). The dipole moment of symmetric diorganozinc reagents can be seen as zero in these linear complexes, which explains their solubility in nonpolar solvents like cyclohexane. Unlike other binary metal alkyls, the diorganozinc species show a low affinity for complexation with ethereal solvent. Bonding in R2Zn is described as employing sp-hybridized orbitals on Zn.[2]

These structures cause zinc to have two bonding d-orbitals and three low-lying non-bonding d-orbitals (see non-bonding orbital), which are available for binding. When zinc lacks electron donating ligands it is unable to obtain coordination saturation, which is a consequence of the large atomic radius and low electron deficiency of zinc. Therefore, it is rare for bridging alkyl or aryl groups to occur due to the weak electron deficiency of the zinc atom. Although, it does occur in some cases such as Ph2Zn (Shown below) and which are the organozinc can form metal clusters (see cluster chemistry). When a ligand is added to the zinc atom both the acceptor and donor character of zinc is enhanced allowing for aggregation.[2] Saturated diorganozinc reagents with bridging aryl groups

Synthesis

Several methods exist for the generation of organozinc compounds. Commercially available diorganozinc compounds are , diethylzinc and diphenylzinc. These reagents are expensive and difficult to handle. In one study[8][9] the active is obtained from much cheaper organobromine precursors:

(2.1)

From zinc metal

Frankland's original synthesis of diethylzinc involves the reaction of ethyl iodide with zinc metal. The zinc must be activated to facilitate this reaction. One of such activated form of zinc employed by Frankland is zinc- couple.[5]

0 2 EtI + 2 Zn → Et2Zn + ZnI2 (2.2)

Riecke zinc, produced by in situ reduction of ZnCl2 with , is another activated form of zinc. This form has proven useful for reactions such as Negishi coupling and Fukuyama coupling. Formation of organozinc reagents is facilitated for alkyl or aryl bearing electron-withdrawing substituents, e.g., and .[10][11]

(2.4)

Functional group exchange

The two most common zinc functional group interconversion reactions are with halides and boron, which is catalyzed by copper iodide (CuI) or base. The boron intermediate is synthesized by an initial reaction followed by treatment with diethyl zinc. This synthesis shows the utility of organozinc reagents by displaying high selectivity for the most reactive site in the molecule, as well as creating useful coupling partners.[12]

(2.5)

This group transfer reaction can be used in allylation, or other coupling reactions (such as Negishi coupling).[13]

(2 β-Silyl diorganozinc compounds

One of the major drawbacks of diorganozinc alkylations is that only one of the two alkyl substituents is transferred. This problem can be solved by using Me3SiCH2- (TMSM), which is a non-transferable group.[14]

(2.7)

Transmetallation

Transmetallation is similar to the interconversions displayed above zinc can exchange with other metals such as , , copper, etc. One example of this reaction is the reaction of diphenylmercury with zinc metal to form diphenylzinc and metallic mercury:

HgPh2 + Zn → ZnPh2 + Hg (2.8)

The benefit of transmetalling to zinc it is often more tolerant of other functional groups in the molecule due to the low reactivity which increases selectivity.[15]

In the synthesis of Maoecrystal V, a directed ortho gives the initial aryl-lithium species, which is transmetallated to the desired arylzinc compound. The arylzinc compound is significantly less reactive than the aryl-lithium species and thus better tolerates the functionality in the subsequent coupling with methyl chlorooxaloacetate. Esters are significantly stable against organozinc reagents.[16]

Organozinc can be obtained directly from zinc metal:[17][18]

(2.10)

In this method zinc is activated by 1,2-dibromoethane and . A key ingredient is which quickly forms a soluble adduct with the organozinc compound thus removing it from the metal surface.

Reactions

In many of their reactions organozincs appear as intermediates.

In the Frankland–Duppa reaction (1863) an (ROCOCOOR) reacts with an alkyl halide R'X, zinc and to the α- hydroxycarboxylic esters RR'COHCOOR[19]

Reformatsky reaction

This can be employed to convert α-haloester and or to a β-hydroxyester. Acid is needed to protonate the resulting during work up. The initial step is an oxidative addition of zinc metal into the carbon-halogen bond, thus forming a carbon-zinc . This C-Zn enolate can then rearrange to the -Zinc enolate via coordination. Once this is formed the other carbonyl containing starting material will coordinate in the manner shown below and give the product after protonation.[20] The benefits of the Reformatsky reaction over the conventional protocols is the following:

1. Allows for exceedingly derivatized ketone substrates 2. The ester enolate intermediate can be formed in the presence of enolizable moieties 3. Well suited for intramolecular reactions

Below shows the six-membered transition state of the Zimmerman–Traxler model ( Control, see Aldol reaction), in which R3 is smaller than R4.[21]

The Reformatsky reaction has been employed in numerous total syntheses such as the synthesis of C(16),C(18)-bis-epi-cytochalasin D:[22]

The Reformatsky reaction even allows for with zinc homo-.[23] A modification of the Reformatsky reaction is the Blaise reaction.[21]

Simmons–Smith reaction

The Simmons–Smith reagent is used to prepare cyclopropanes from olefin using methylene iodide as the methylene source. The reaction is effected with zinc. The key zinc-intermediate formed is a (iodomethyl)zinc iodide which reacts with to afford the cyclopropanated product. The rate of forming the active zinc species is increased via ultrasonication since the initial reaction occurs at the surface of the metal.

(3.4)

(3.5)

Although the mechanism has not been fully elaborated it is hypothesized that the organozinc intermediate is a metal-carbenoid. The intermediate is believed to be a three-centered "butterfly-type". This intermediate can be directed by substituents, such as , to deliver the cyclopropane on the same side of the molecule. Zinc- copper couple is commonly used to activate zinc.[21]

(3.6)

Titanium–zinc methylenation

Organozinc compounds derived from methylene bromide or iodide can electrophilically add to carbonyl groups to form terminal alkenes.[24] The reaction is [25] mechanistically related to the Tebbe reaction and can be catalyzed by various Lewis acids (e.g. TiCl4 or Al2Me6). The reaction is used to introduce deuterium into molecules for or as an alternative to the .

Negishi coupling

This powerful carbon-carbon bond forming cross-coupling reactions combines an organic halide and an organozinc halide reagent in the presence of a nickel or catalyst. The organic halide reactant can be alkenyl, aryl, allyl, or propargyl. Alkylzinc coupling with alkyl halides such as bromides and chlorides have also been reported with active catalysts such as Pd-PEPPSI precatalysts, which strongly resist beta-hydride elimination (a common occurrence with alkyl substituents).[26] Either diorganic species or organozinc halides can be used as coupling partners during the transmetallation step in this reaction. Despite the low reactivity of organozinc reagents on organic , these reagents are among the most powerful metal toward palladium.[27]

2− Alkylzinc species require the presence of at least a stoichiometric amount of halide in solution to form a "zincate" species of the form RZnX3 , before it can undergo to the palladium centre.[28] This behavior contrasts greatly with the case of aryl zinc species. A key step in the catalytic cycle is a transmetalation in which a zinc halide exchanges its organic substituent for another halogen with the metal center.

An elegant example of Negishi coupling is Furstner's synthesis of amphidinolide T1:[29]

Fukuyama coupling

Fukuyama coupling is a palladium-catalyzed reaction involving the coupling of an aryl, alkyl, allyl, or α,β- unsaturated compound. This thioester compound can be coupled to a wide range of organozinc reagents in order to reveal the corresponding ketone product. This protocol is useful due to its sensitivity to functional groups such as ketone, , aromatic halides, and even . The chemoselectivity observed indicates ketone formation is more facile than oxidative addition of palladium into these other moieties.[30]

(3.8)

A further example of this coupling method is the synthesis of (+)-biotin. In this case, the Fukuyama coupling takes place with the thiolactone:[31]

Barbier reaction

The Barbier reaction involves of a carbanion equivalent to a carbonyl. The conversion is similar to the . The organozinc reagent is generated via an oxidative addition into the alkyl halide. The reaction produces a primary, secondary, or tertiary via a 1,2-addition. The Barbier reaction is advantageous because it is a one-pot process: the organozinc reagent is generated in the presence of the carbonyl substrate. Organozinc reagents are also less water sensitive, thus this reaction can be conducted in water. Similar to the Grignard reaction, a Schlenk equilibrium applies, in which the more reactive dialkylzinc can be formed.[21]

(3.10)

The mechanism resembles the Grignard reaction, in which the metal alkoxide can be generated by a stepwise pathway, through single electron transfer, or concerted reaction pathway via a cyclic transition state. An example of this reaction is in Danishefsky’s synthesis of cycloproparadicicol. By using the organozinc conditions the other functionality of the dienone and the are tolerated:[32]

.

Zinc acetylides

The formation of the zinc proceeds via the intermediacy of a dialkynyl zinc (functional group exchange). Catalytic processes have been developed such as Merck’s process.[33] Propargylic alcohols can be synthesized from zinc acetylides. These versatile intermediates can then be used for a wide range of chemical transformations such as cross-coupling reactions, , and pericyclic reactions.[34]

(3.12)

In the absence of ligands, the reaction is slow and inefficient. In the presence of chiral ligands, the reaction is fast and gives high conversion. Noyori determined that a monozinc-ligand complex is the active species.[35]

(3 Diastereoselectivity for addition of organozinc reagents into aldehydes can be predicted by the following model by Noyori and David A. Evans:[36]

(3

The α- of the ligand dictates observed of the propargylic alcohols The between the aldehyde substituent and the ligand are less important but still dictate the favored conformation

Zinc-acetylides are used in the HIV-1 reverse transcriptase inhibitor as well as in Merck’s ephedrine derivatives .[37]

(3.15)

Organozincates

The first organozinc (organozincate) was reported in 1858 by James Alfred Wanklyn,[38] an assistant to Frankland and concerned the reaction of elemental with diethylzinc:

2 Na + 3 ZnEt2 → 2 NaZnEt3 + Zn (4.1)

Organozinc compounds that are strongly Lewis acidic are vulnerable to nucleophilic attack by alkali metals, such as sodium, and thus form these ‘ate compounds’. Two types of organozincates are recognized: tetraorganozincates ([R4Zn]M2), which are dianionic, and triorganozincates ([R3Zn]M), which are monoanionic. Their structures, which are determined by the ligands, have been extensively characterized.[3]

Synthesis

Tetraorganozincates such as [Me4Zn]Li2 can be formed by mixing Me2Zn and MeLi in a 1:2 molar ratio of the rectants. Another example synthetic route to forming spriocyclic organozincates is shown below:[3]

Triorganozincates compounds are formed by treating a diorganozinc such as (Me3SiCH2)2Zn with an (K), or an alkali earth metal (Ba, Sr, or Ca). One [39] example is [(Me3SiCH2)3Zn]K. Triethylzincate degrades to sodium hydridoethylzincate(II) as a result of beta-hydride elimination:

2 NaZnEt3 → Na2Zn2H2Et4 + 2 C2H4 (4.3)

The product is an edge-shared bitetrahedral structure, with bridging hydride ligands.

Reactions

Although less commonly studied, organozincates often have increased reactivity and selectivity compared to the neutral diorganozinc compounds. They have been useful in stereoselective alkylations of and related carbonyls, ring opening reactions. Aryltrimethylzincates participate in vanadium mediated C-C forming reactions.[3]

(4.4)

Organozinc(I) compounds

Low valent organozinc compounds having a Zn–Zn bond are also known. The first such compound, , was reported in 2004.[40]

See also Compounds of zinc

References 1. Knochel, Paul; Millot, Nicolas; Rodriguez, Alain L.; Tucker, Charles 16. Lu, Ping; Gu, Zhenhua; Zakarian, Armen (2013). "Total Synthesis of E. (2004). Organic Reactions. doi:10.1002/0471264180.or058.02 (htt Maoecrystal V: Early-Stage C–H Functionalization and Lactone ps://doi.org/10.1002%2F0471264180.or058.02). ISBN 0471264180. Assembly by Radical Cyclization" (https://www.ncbi.nlm.nih.gov/pmc/ 2. The Chemistry of Organozinc Compounds (http://www.wiley.com/Wil articles/PMC4118676). Journal of the American . eyCDA/WileyTitle/productCd-0470093374.html) (Patai Series: The 135 (39): 14552–5. doi:10.1021/ja408231t (https://doi.org/10.1021%2 Chemistry of Functional Groups (http://www.wiley.com/WileyCDA/Se Fja408231t). PMC 4118676 (https://www.ncbi.nlm.nih.gov/pmc/article ction/id-11128.html)), (Eds. Z. Rappoport and I. Marek), John Wiley & s/PMC4118676). PMID 24047444 (https://pubmed.ncbi.nlm.nih.gov/2 Sons: Chichester, UK, 2006, ISBN 0-470-09337-4. 4047444). 3. Organozinc reagents – A Practical Approach, (Eds. P. Knochel and P. 17. Krasovskiy, Arkady; Malakhov, Vladimir; Gavryushin, Andrei; Jones), Oxford Medical Publications, Oxford, 1999, ISBN 0-19- Knochel, Paul (2006). "Efficient Synthesis of Functionalized 850121-8. Organozinc Compounds by the Direct Insertion of Zinc into Organic 4. Synthetic Methods of Organometallic and Inorganic Chemistry Vol 5, Iodides and Bromides". Angewandte Chemie International Edition. 45 (36): 6040–6044. doi:10.1002/anie.200601450 (https://doi.org/10. Copper, Silver, Gold, Zinc, , and Mercury, W.A. Herrmann Ed., ISBN 3-13-103061-5 1002%2Fanie.200601450). PMID 16900548 (https://pubmed.ncbi.nl m.nih.gov/16900548). 5. E. Frankland, Liebigs Ann. Chem.,1849, 71, 171 18. In this example the arylzinc iodide continues to react with allyl 6. Elschenbroich, C. ”Organometallics” (2006) Wiley-VCH: Weinheim. bromide in a nucleophilic displacement ISBN 978-3-527-29390-2 19. Merck Index (https://archive.org/details/merckindexency00onei). 7. John Bacsa; Felix Hanke; Sarah Hindley; Rajesh Odedra; George R. Merck & Co. Frankland–Duppa reaction. Darling; Anthony C. Jones; Alexander Steiner (2011). "The Solid State Structures of Dimethylzinc and Diethylzinc" (https://www.ncbi.nl 20. Fürstner, Alois (1989). "Recent Advancements in the Reformatsky Reaction". Synthesis. 1989 (8): 571–590. doi:10.1055/s-1989-27326 m.nih.gov/pmc/articles/PMC3326375). Angewandte Chemie International Edition. 50 (49): 11685–11687. (https://doi.org/10.1055%2Fs-1989-27326).) doi:10.1002/anie.201105099 (https://doi.org/10.1002%2Fanie.20110 21. Kurti, L.; Czako, B. ‘‘Strategic Applications of Named Reactions in 5099). PMC 3326375 (https://www.ncbi.nlm.nih.gov/pmc/articles/PM Organic Synthesis’’; Elsevier: Burlington, 2005. C3326375). PMID 21919175 (https://pubmed.ncbi.nlm.nih.gov/21919 22. Vedejs, E.; Duncan, S. M. (2000). "A Synthesis of C(16),C(18)-Bis- 175). epi-cytochalasin D via Reformatsky Cyclization". The Journal of 8. Kim, Jeung Gon; Walsh, Patrick J. (2006). "From Aryl Bromides to Organic Chemistry. 65 (19): 6073–81. doi:10.1021/jo000533q (https:// Enantioenriched Benzylic Alcohols in a Single Flask: Catalytic doi.org/10.1021%2Fjo000533q). PMID 10987942 (https://pubmed.nc Asymmetric Arylation of Aldehydes". Angewandte Chemie bi.nlm.nih.gov/10987942). International Edition. 45 (25): 4175–4178. 23. Kumwaijima,I.; et al. J. Am. Chem. 1987, 109, 8056 doi:10.1002/anie.200600741 (https://doi.org/10.1002%2Fanie.20060 24. Takai, Kazuhiko; Hotta, Yuji; Oshima, Koichiro; Nozaki, Hitosi (1980). 0741). PMID 16721894 (https://pubmed.ncbi.nlm.nih.gov/16721894). "Wittig-type Reaction of Dimetallated Carbodianion Species as 9. In this one-pot reaction bromobenzene is converted to phenyllithium Produced by Zinc Reduction of gem-Polyhalogen Compounds in the by reaction with 4 equivalents of n-butyllithium, then transmetalation Presence of Lewis Acids" (http://www.journal.csj.jp/doi/pdf/10.1246/b with forms diphenylzinc which continues to react in an csj.53.1698). Bulletin of the Chemical Society of Japan. 53 (6): 1698– asymmetric reaction first with the MIB ligand and then with 2- 1702. doi:10.1246/bcsj.53.1698 (https://doi.org/10.1246%2Fbcsj.53.1 naphthylaldehyde to the alcohol. In this reaction formation of 698). diphenylzinc is accompanied by that of lithium chloride, which 25. Trost, Barry; Fleming, Ian; Schreiber, Stuart (1991). "Transformation unchecked, catalyses the reaction without MIB involvement to the of the into Nonhydroxylic Groups". Comprehensive racemic alcohol. The salt is effectively removed by chelation with Organic Synthesis Volume 1: Additions to CX π-Bonds, Part 1 (1st tetraethylethylene diamine (TEEDA) resulting in an enantiomeric ed.). Newyork: Pergamon Press. pp. 749–751. doi:10.1016/B978-0- excess of 92%. 08-052349-1.00020-2 (https://doi.org/10.1016%2FB978-0-08-052349 10. Rieke, R. D. (1989). "Preparation of Organometallic Compounds from -1.00020-2). ISBN 9780080405926. Highly Reactive Metal Powders". Science. 246 (4935): 1260–1264. 26. S. Sase, M. Jaric, A. Metzger, V. Malakhov, P. Knochel, J. Org. Bibcode:1989Sci...246.1260R (https://ui.adsabs.harvard.edu/abs/198 Chem., 2008, 73, 7380-7382 9Sci...246.1260R). doi:10.1126/science.246.4935.1260 (https://doi.or 27. Nicolaou, K. C.; Bulger, Paul G.; Sarlah, David (2005). "Palladium- g/10.1126%2Fscience.246.4935.1260). PMID 17832221 (https://pub Catalyzed Cross-Coupling Reactions in Total Synthesis". med.ncbi.nlm.nih.gov/17832221). Angewandte Chemie International Edition. 44 (29): 4442–4489. 11. Negishi, Ei-Ichi (2002). "A genealogy of Pd-catalyzed cross- doi:10.1002/anie.200500368 (https://doi.org/10.1002%2Fanie.20050 coupling". Journal of . 653 (1–2): 34–40. 0368). PMID 15991198 (https://pubmed.ncbi.nlm.nih.gov/15991198).) doi:10.1016/S0022-328X(02)01273-1 (https://doi.org/10.1016%2FS0 28. McCann, L. C.; Hunter, H. N.; Cyburne, J. A. C.; Organ, M. G (2012). 022-328X%2802%2901273-1). "Higher-Order Zincates as Transmetalators in Alkyl-Alkyl Negishi 12. Langer, Falk; Schwink, Lothar; Devasagayaraj, Arokiasamy; Cross-Coupling". Angew. Chem. Int. Ed. 51 (28): 7024–7027. Chavant, Pierre-Yves; Knochel, Paul (1996). "Preparation of doi:10.1002/anie.201203547 (https://doi.org/10.1002%2Fanie.20120 Functionalized Dialkylzincs via a Boron−Zinc Exchange. Reactivity 3547). PMID 22685029 (https://pubmed.ncbi.nlm.nih.gov/22685029). and Catalytic Asymmetric Addition to Aldehydes". The Journal of 29. Aïssa, Christophe; Riveiros, Ricardo; Ragot, Jacques; Fürstner, Alois Organic Chemistry. 61 (23): 8229–8243. doi:10.1021/jo961129n (http (2003). "Total Syntheses of Amphidinolide T1, T3, T4, and T5". s://doi.org/10.1021%2Fjo961129n). ISSN 0022-3263 (https://www.wo Journal of the American Chemical Society. 125 (50): 15512–20. rldcat.org/issn/0022-3263). PMID 11667810 (https://pubmed.ncbi.nl doi:10.1021/ja038216z (https://doi.org/10.1021%2Fja038216z). m.nih.gov/11667810). PMID 14664598 (https://pubmed.ncbi.nlm.nih.gov/14664598). 13. Naka,H; et al.New J. Chem., 2010, 34, 1700–1706 30. Tokuyama, Hidetoshi; Yokoshima, Satoshi; Yamashita, Tohru; 14. Knochel,P.; et al. Angel. Chem. Int. Ed. Engl. 1997, volume 36, 1496- Fukuyama, Tohru (1998). "A novel ketone synthesis by a palladium- 1498 catalyzed reaction of thiol esters and organozinc reagents". 15. Markies, P; Schat, Gerrit; Akkerman, Otto S.; Bickelhaupt, F.; Spek, Tetrahedron Letters. 39 (20): 3189–3192. doi:10.1016/S0040- Anthony L. (1992). "Complexation of diphenylzinc with simple . 4039(98)00456-0 (https://doi.org/10.1016%2FS0040-4039%2898%2 Crystal structures of the complexes Ph2Zn·glyme and 900456-0). Ph2Zn·diglyme". J. Organomet. Chem. 430: 1–13. doi:10.1016/0022- 31. Shimizu, Toshiaki; Seki, Masahiko (2000). "Facile synthesis of (+)- 328X(92)80090-K (https://doi.org/10.1016%2F0022-328X%2892%29 biotin via Fukuyama coupling reaction". Tetrahedron Letters. 41 (26): 80090-K). 5099–5101. doi:10.1016/S0040-4039(00)00781-4 (https://doi.org/10. 1016%2FS0040-4039%2800%2900781-4). 32. Yang, Zhi-Qiang; Geng, Xudong; Solit, David; Pratilas, Christine A.; 36. Evans, D. (1988). "Stereoselective organic reactions: Catalysts for Rosen, Neal; Danishefsky, Samuel J. (2004). "New Efficient carbonyl addition processes". Science. 240 (4851): 420–6. Synthesis of Resorcinylic Macrolides via Ynolides: Establishment of Bibcode:1988Sci...240..420E (https://ui.adsabs.harvard.edu/abs/198 Cycloproparadicicol as Synthetically Feasible Preclinical Anticancer 8Sci...240..420E). doi:10.1126/science.3358127 (https://doi.org/10.11 Agent Based on Hsp90 as the Target". Journal of the American 26%2Fscience.3358127). PMID 3358127 (https://pubmed.ncbi.nlm.ni Chemical Society. 126 (25): 7881–9. doi:10.1021/ja0484348 (https:// h.gov/3358127). doi.org/10.1021%2Fja0484348). PMID 15212536 (https://pubmed.nc 37. Thompson, A. S.; Corley, E. G.; Huntington, M. F.; Grabowski, E. J. J. bi.nlm.nih.gov/15212536). Tetrahedron Lett. 1995, 36, 8937-8940 33. Li, Z.; Upadhyay, V.; DeCamp, A. E.; DiMichele, L.; Reider, P. J. 38. J. A. Wanklyn (1858). "Ueber einige neue Aethylverbindungen, Synthesis 1999, 1453-1458. welche Alkalimetalle enthalten" (https://zenodo.org/record/1427111). 34. Soai, Kenso; Niwa, Seiji (1992). "Enantioselective addition of Liebigs Annalen. 108 (67): 67–79. doi:10.1002/jlac.18581080116 (htt organozinc reagents to aldehydes". Chemical Reviews. 92 (5): 833– ps://doi.org/10.1002%2Fjlac.18581080116). 856. doi:10.1021/cr00013a004 (https://doi.org/10.1021%2Fcr00013a 39. Lennartson, Anders; Håkansson, Mikael; Jagner, Susan (2007). 004). "Facile Synthesis of Well-Defined Sodium Hydridoalkylzincates(II)". 35. Noyori, Ryoji; Kitamura, Masato (1991). "Enantioselective Addition of Angewandte Chemie International Edition. 46 (35): 6678–6680. Organometallic Reagents to Carbonyl Compounds: Chirality doi:10.1002/anie.200701477 (https://doi.org/10.1002%2Fanie.20070 Transfer, Multiplication, and Amplification". Angewandte Chemie 1477). PMID 17665387 (https://pubmed.ncbi.nlm.nih.gov/17665387). International Edition in English. 30: 49–69. 40. Schulz, Stephan (2010). "Low-Valent Organometallics-Synthesis, doi:10.1002/anie.199100491 (https://doi.org/10.1002%2Fanie.19910 Reactivity, and Potential Applications". Chemistry: A European 0491). Journal. 16 (22): 6416–28. doi:10.1002/chem.201000580 (https://doi. org/10.1002%2Fchem.201000580). PMID 20486240 (https://pubmed. ncbi.nlm.nih.gov/20486240).

External links

Zinc in organic synthesis (http://www.chem.wisc.edu/areas/reich/orgmet/zinc.htm) Organozinc Compounds produced by BASF Corporation (http://www.rznx.basf.com)

Retrieved from "https://en.wikipedia.org/w/index.php?title=Organozinc_compound&oldid=951424339"

This page was last edited on 17 April 2020, at 02:48 (UTC).

Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply. By using this site, you agree to the Terms of Use and Privacy Policy. Wikipedia® is a registered trademark of the Wikimedia Foundation, Inc., a non-profit organization.